The HIPPO Transducer YAP and Its Targets CTGF and Cyr61 Drive a Paracrine Signalling in Cold Atmospheric Plasma-Mediated Wound Healing

Page created by Jonathan Walker
 
CONTINUE READING
The HIPPO Transducer YAP and Its Targets CTGF and Cyr61 Drive a Paracrine Signalling in Cold Atmospheric Plasma-Mediated Wound Healing
Hindawi
Oxidative Medicine and Cellular Longevity
Volume 2020, Article ID 4910280, 14 pages
https://doi.org/10.1155/2020/4910280

Research Article
The HIPPO Transducer YAP and Its Targets CTGF and
Cyr61 Drive a Paracrine Signalling in Cold Atmospheric
Plasma-Mediated Wound Healing

 Debarati Shome ,1 Thomas von Woedtke,1 Katharina Riedel,2 and Kai Masur 1

 1
 Leibniz Institute for Plasma Science and Technology, A Member of the Leibniz Research Alliance Leibniz Health Technology, Felix-
 Hausdorff-Straße 2, Greifswald 17489, Germany
 2
 Institute for Microbiology, University of Greifswald, Felix-Hausdorff-Straße 8, Greifswald 17489, Germany

 Correspondence should be addressed to Kai Masur; kai.masur@inp-greifswald.de

 Received 9 September 2019; Revised 27 January 2020; Accepted 29 January 2020; Published 13 February 2020

 Academic Editor: Reiko Matsui

 Copyright © 2020 Debarati Shome et al. This is an open access article distributed under the Creative Commons Attribution License,
 which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

 Reactive species play a pivotal role in orchestrating wound healing responses. They act as secondary messengers and drive
 redox-signalling pathways that are involved in the homeostatic, inflammatory, proliferative, and remodelling phases of wound
 healing. The application of Cold Atmospheric Plasma (CAP) to the wound site produces a profusion of short- and long-lived
 reactive species that have been demonstrated to be effective in promoting wound healing; however, knowledge of the
 mechanisms underlying CAP-mediated wound healing remains scarce. To address this, an in vitro coculture model was used to
 study the effects of CAP on wound healing and on paracrine crosstalk between dermal keratinocytes and fibroblasts. Using this
 coculture model, we observed a stimulatory effect on the migration ability of HaCaT cells that were cocultured with dermal
 fibroblasts. Additionally, CAP treatment resulted in an upregulation of the HIPPO transcription factor YAP in HaCaTs and
 fibroblasts. Downstream effectors of the HIPPO signalling pathway (CTGF and Cyr61) were also upregulated in dermal
 fibroblasts, and the administration of antioxidants could inhibit CAP-mediated wound healing and abrogate the gene expression
 of the HIPPO downstream effectors. Interestingly, we observed that HaCaT cells exhibited an improved cell migration rate
 when incubated with CAP-treated fibroblast-conditioned media compared to that observed after incubation with untreated
 media. An induction of CTGF and Cyr61 secretion was also observed upon CAP treatment in the fibroblast-conditioned media.
 Finally, exposure to recombinant CTGF and Cyr61 could also significantly improve HaCaT cell migration. In summary, our
 results validated that CAP activates a regenerative signalling pathway at the onset of wound healing. Additionally, CAP also
 stimulated a reciprocal communication between dermal fibroblasts and keratinocytes, resulting in improved keratinocyte wound
 healing in coculture.

1. Introduction The short- and long-lived ROS and RNS produced by CAP
 are involved in redox signalling in vitro and in vivo. Wound
Cold Atmospheric Plasma (CAP) is an emerging tool of healing involves a coordinated infiltration of dermal cells that
biomedical and clinical importance that controls cellular is followed by infiltration of immune cells, extracellular
processes such as wound healing, immunomodulation, cell matrix deposition, and reepithelization over distinct yet
death, and cancer. CAP is a partially ionized molecular or interdependent phases that are regulated by a large number
noble gas or gas mixture that generates a plethora of reactive of transcription factors or transcriptional events [2–5]. ROS
oxygen and nitrogen species (ROS/RNS), radiation (ultravio- are critical secondary messengers that orchestrate wound
let (UV), visible, and infrared), and small chemical entities healing processes by regulating the recruitment of immune
such as neutral molecules, ions, electrons, and exited atoms cells, angiogenesis, and optimal perfusion of blood at the
under atmospheric pressure and ambient temperature [1]. wound site [6]. Recent reports also indicate that CAP
2 Oxidative Medicine and Cellular Longevity

supports wound healing by influencing cell viability [7, 8], tion of organ survival and regenerative pathways further
proliferation [9], migration [10], inflammation [11], dermal establishes the usefulness of CAP in redox-mediated wound
regeneration, and reepithelization [12]. Currently, only a small management and repair.
number of studies have examined the cellular and molecular
effects of CAP-mediated wound healing [1, 13, 14] and 2. Materials and Methods
the mechanisms underlying CAP-mediated wound healing
in the context of regenerative pathways require further 2.1. Cell Culture. Human skin keratinocytes (HaCaT) and
investigation. the immortalized skin fibroblast cell line GM00637 (GM
 The HIPPO signalling pathway was initially identified Fbs) were cultured in Roswell Park Memorial Institute
in Drosophila melanogaster, and this pathway regulates 1640 cell culture medium (RPMI 1640, Invitrogen) supple-
cellular proliferation, organ size, and survival and is highly mented with 10% fetal calf serum (FCS, Invitrogen),
conserved in mammals [15–17]. The core HIPPO pathway 0.1 mg/ml penicillin/streptomycin, and 2 mM L-glutamine
in vertebrates consists of a kinase cascade where one of (PAN Biotech, Germany). For 2D scratch assays, the indi-
the major transcriptional coactivators is Yes-associated cated number of cells (either in mono- or coculture), as per
protein (YAP) [18, 19]. Nonphosphorylated YAP translo- experimental requirements, was incubated in 96-well image
cates to the nucleus and induces target gene transcription lock culture plates (Essen BioScience) in a starved cell culture
by interacting with the transcription factors TEA domain medium (supplemented with 1% FCS) 16-18 hr prior to their
family member proteins (TEAD 1-4) [20]. In particular, experimental use.
YAP-mediated target gene transcription depends upon its
cellular localization (cytoplasmic or nuclear) [21–24]. 2.1.1. Coculture. Direct coculture of HaCaTs and GM Fbs
 The HIPPO signalling pathway is of paramount impor- was used for cell migration assays and metabolic assays in
tance in the oxidative stress response and wound healing. 96-well image lock culture plates (Essen BioScience) or nor-
YAP is involved in redox homeostasis regulation, and mal 96-well plates (Sarstedt), respectively. For coculture
its deletion can lead to redox sensitivity [22]. Certain experiments, GM Fbs were seeded first into a 96-well plate,
YAP-targeting genes encode secretory proteins and may pos- and once they adhered to the plate, the indicated number of
sess regenerative potential [25]. The downstream target genes HaCaTs was seeded on top of the fibroblasts in a manner that
of YAP that are of primary interest in this study are Connec- was dependent upon experimental type. The exact cell num-
tive Tissue Growth Factor (CTGF) and Cysteine-rich angio- bers are described in the respective method sections. The cell
genic protein 61 (Cyr61), members of CCN family proteins. numbers (in mono- or coculture) and the experimental type
According to literature, YAP moves to the nucleus and inter- are also summarized in Figure S1.
acts with TEAD to promote the transcription of these matri-
cellular proteins that are involved in cell adhesion and 2.2. Harvest of GM-Fb-Conditioned Medium (GM-Fb-CM).
migration [26, 27]. These extracellular matrix-associated To harvest the conditioned medium (CM), GM Fbs
signalling proteins are structurally related heparin-binding (0:1 × 106 /well) were seeded onto a 24-well plate in complete
proteins and are synthesized during cutaneous wound heal- RPMI 1640 (supplemented with 10% FCS) until the cells
ing in dermal fibroblasts [28]. Due to its ability to modulate were approximately 80% confluent. The medium was then
the activities of several growth factors and cytokines, CTGF replaced with 1% FCS medium for 16-18 hr. Next, fibroblasts
possesses the potential to regulate diverse biological pro- were incubated with untreated or CAP-treated medium
cesses including cell adhesion, migration [29], proliferation for another 18 hr. Following the incubation, conditioned
[30], angiogenesis [31], and early wound healing and repair medium was collected. The CM was centrifuged at 2000 rpm
[32]. A physiological role of Cyr61 in regulating inflamma- for 5 min, filtered through a 0.22 μm syringe filter, and kept
tion, adhesion, angiogenesis, proliferation, matrix remodel- at 4°C for immediate use. For longer storage, the medium
ling, and cell-ECM interaction during wound repair has was stored at −80°C.
also been identified [26, 28, 33, 34]. Oxidative stress plays a
role in the alteration of CTGF expression in fibroblasts 2.3. CAP Treatment. CAP treatment was performed using
[35]. Qin et al. found that Cyr61 expression is induced by an atmospheric pressure argon plasma jet kINPen MED
oxidative stress and that a positive feedback loop continues (neoplas tools, Greifswald, Germany). The electrical safety
to generate ROS and upregulate Cyr61 in fibroblasts [36]. of the kINPen was certified and complies with EU standards
Taken together, these reports strongly suggest the presence [37]. This is the first jet worldwide to be accredited as a med-
of a redox-induced HIPPO signalling axis in wound healing ical device class IIa according to European Council Directive
that is complex and poorly understood. 93/42/EEC for use in wound treatment, and this device lacks
 In this study, for the first time, we demonstrate the acti- mutagenic and genotoxic effects [37–39]. The jet was oper-
vation of the YAP-CTGF-Cyr61 axis in response to CAP ated at a voltage of 2-6 kVpp and a frequency of 1 MHz.
treatment in a well-characterized human epithelial cell line The gas flow rate was set at 5 SLM (standard litres per
(HaCaT) and in a human nonneonatal skin fibroblast cell minute) to treat 5 ml of RPMI 1640 supplemented with 1%
line (GM00637). CAP-mediated acceleration of wound heal- FCS in a 60 mm dish for 10, 30, and 60 s. CAP-treated media
ing indicates the presence of a paracrine signalling pathway were immediately used to treat the cells for further experi-
that is driven by the extracellular matrix-associated signalling ments. Cells incubated with untreated RPMI medium served
proteins CTGF and Cyr61 in a coculture model. The activa- as negative controls for all experiments.
Oxidative Medicine and Cellular Longevity 3

2.4. Metabolic Activity Assay. To determine metabolic Fisher Scientific), and qPCR was performed in triplicate with
activity, 2:1 × 104 cells (in monoculture) were seeded into a SYBR Green I Master mix (Roche Diagnostics, Germany)
96-well plate (Sarstedt) 24 hours prior to the start of the specific for the genes YAP, CTGF, and Cyr61. Primer
experiment. For coculture, 0:7 × 104 GM Fbs and 1:4 × 104 sequences are provided in Fig S2. The housekeeping gene
HaCaTs were seeded onto the plate. After incubation with β-actin was used as the internal control for normalization.
CAP-treated media for 3 and 24 hr, the wells were loaded Gene expression analysis was performed using the ΔΔCT
with 100 μM of resazurin (Alfa Aesar) that is transformed method [40]. The final fold change in gene expression upon
to fluorescent resorufin by metabolically active cells. The plasma treatment was determined according to the ratio
plate was incubated for 2 h at 37°C, and fluorescence was of expression in the respective sample corresponding to
measured using a Tecan plate reader (Infinite M200 Pro) at the control.
λex 535 nm and λem 590 nm. The fluorescence values were
normalized to those of the untreated controls. 2.6. Immunoblotting. HaCaTs and GM Fbs were seeded at a
 density of 0:5 × 106 cells in a 6-well plate. After the cells
2.5. Cell Migration Assay. The cell migratory behaviour of reached confluency, they were serum starved for 16-18 hours,
HaCaTs and GM Fbs was assayed using a cell migration assay scratches were created using a 10 μl pipette tip, and the cells
that incorporated an Essen BioScience wound maker. For were subsequently incubated with CAP-treated medium or
this assay, separate monocultures of HaCaT and GM Fbs untreated medium. Cells were harvested into ice-cold PBS
and coculture of HaCaT and GM Fbs were used. For the after 3, 6, and 24 hr. Next, the samples were lysed in RIPA
monocultures, either 6 × 104 HaCaTs/well or GM Fbs/well buffer that was supplemented with protease and phosphatase
were seeded into Essen BioScience 96-well image lock plates. inhibitors (Roche) for 20 min on ice. Lysates were then cen-
For the coculture, 2 × 104 GM Fbs/well were seeded into trifuged at 15000 g for 15 min at 4°C, and total extracted pro-
image lock plates and incubated until the fibroblasts attached tein was quantified using the DC protein assay (Bio-Rad).
to the plates, and then 4 × 104 HaCaTs/well were seeded onto Total protein (25 μg) was resolved by 10% SDS-PAGE
the top of the fibroblasts at a ratio of 1 : 2 fibroblasts: kerati- (Bio-Rad) and then transferred and blotted on PVDF
nocytes within the coculture. Several other ratios (1 : 3, 1 : 5) membranes. The membranes were probed with antitotal
of HaCaTs and GM Fbs were also tested for the cell migration YAP (Santa Cruz Biotechnology), phospho-YAP (CST),
assays (data not shown), but in further experiments, a 1 : 2 CTGF (Novex Biotechnology), Cyr61 (CST), and β-actin
ratio of HaCaT and GM Fbs was used. After complete adher- (CST) primary antibodies, and this was followed by incuba-
ence of HaCaTs and GM-Fbs in mono- and cocultures, the tion with secondary horse radish peroxide- (HRP-) coupled
cells were starved using 1% FCS for 16-18 hr prior to the antibodies (CST) and subsequent detection of the bands
experiment. The Essen BioScience wound maker was used to using chemiluminescence (Applied Biosystems). Band inten-
create scratches, and the cells were then incubated with either sities were quantified with Image Studio Lite (version 5.2),
10 or 30 s plasma-treated medium or untreated medium normalized to housekeeping β-actin, and expressed as a fold
as indicated. change compared to corresponding untreated control.
 For other experiments, recombinant CTGF (10-50 ng/ml)
 2.7. N-Acetyl Cysteine (NAC) Treatment. A 100 mM stock
and Cyr61 (100-1000 ng/ml) (PeproTech GmBH) and
 solution was prepared by dissolving NAC powder (Sigma-
GM-Fb-CMs were used on HaCaT monocultures seeded at
 Aldrich) into 1x sterile PBS. This solution was sterilized using
a density of 4 × 104 cells/well. a 0.2 μM sterile membrane filter, and aliquots of the stock
 The cell migration rate was monitored using an IncuCyte solution were stored at −20°C. HaCaTs and GM Fbs were
S3 live cell analysis system for up to 24 hr. The relative wound incubated with 2.5 mM NAC in the presence or absence of
density was measured with the in-built software of the CAP-treated media for the designated times and then
IncuCyte S3 2018A. The relative wound density (%) is the harvested for further experiments.
ratio of the occupied area of the initially scratched area with
respect to the total scratched area. The graphs are plotted as a 2.8. Enzyme-Linked Immunosorbent Assay (ELISA). GM-Fb-
function of relative wound density (%) in y-axis in compari- CM was collected 18 hours after incubation with 10 and 30 s
son to the time at the x-axis. of CAP-treated or untreated medium and analysed using
 enzyme-linked immunosorbent assay (ELISA). The concen-
2.5.1. mRNA Expression Analysis. Quantitative real-time trations of secreted CTGF (PeproTech GmBH, catalogue
PCR was performed using a 96-well LightCycler 480 qPCR no: 900-K317) and Cyr61 (R&D systems, catalogue no:
system according to the manufacturer’s protocol (Roche DCYR10) were quantitatively determined according to the
Diagnostics, Germany). Briefly, 0:5 × 106 HaCaTs or GM manufacturer’s protocol. ELISA values were measured in
Fbs were seeded into a 6-well plate, and once the monocul- duplicate from two different biological experiments for each
tures reached confluence, the cells were starved for 16-18 sample to minimize the intra- and interassay variability.
hours, scratches were made using a 10 μl pipette tip, and
the cells were subsequently incubated with CAP-treated 2.9. Statistical Analysis. Graphics and statistics were per-
medium or untreated control medium. Total RNA was iso- formed using prism 8.02 (GraphPad PRISM software,
lated with an RNA mini kit (Bio & Sell, Germany) after 3, San Diego, USA). Mean and standard deviation were cal-
6, 18, and 24 hr. 1 μg of RNA was transcribed to cDNA using culated and analysed according to unpaired t-test with
a High Capacity cDNA Reverse Transcription Kit (Thermo Welch’s correction and ordinary one-way or 2-way analysis
4 Oxidative Medicine and Cellular Longevity

 HaCaT GM Fbs Coculture
 150 150
 150
 % metabolic activity

 % metabolic activity

 % metabolic activity
 100 ⁎ 100 ⁎⁎ ⁎
 100

 50 50 50

 0 0 0
 con 10 s 30 s 60 s con 10 s 30 s 60 s con 10 s 30 s 60 s
 (a)
 HaCaT GM Fbs Coculture
 150 150 150
 % metabolic activity

 % metabolic activity
 % metabolic activity

 100 100 100
 ⁎⁎⁎
 ⁎⁎⁎⁎

 50 50 50

 0 0 0
 con 10 s 30 s 60 s con 10 s 30 s 60 s con 10s 30 s 60 s
 (b)

Figure 1: Influence of CAP on the metabolic activity in mono- and coculture: the metabolic activity of HaCaTs, GM Fbs, and coculture after
3 hr in response to 10, 30, and 60 s of CAP treatments. Untreated RPMI served as control (a). Metabolic activity of the same cell types after 24
hours of CAP treatment (b). Data are presented as mean ± SD of three independent experiments performed in triplicate. Statistical
comparisons were performed using 1-way ANOVA.

of variance (ANOVA). A p value < 0.05 was considered results, CAP did not significantly alter relative wound density
statistically significant (∗ p ≤ 0:05, ∗∗ p ≤ 0:01, ∗∗∗ p ≤ 0:001, in either of the cell lines after 10 and 30 s of treatments
and ∗∗∗∗ p ≤ 0:0001). (Figures 2(a) and 2(b), first and second column); however,
 CAP significantly accelerated relative wound density in cocul-
3. Results ture after 12 hours (Figures 2(a) and 2(b), third column).
 Wound closure images of both the monocultures and the
3.1. Influence of CAP on Metabolic Activity in Mono- and coculture are provided in the supplementary figures (S3, S4,
Coculture. There are compelling evidences within the litera- and S5). We also compared the relative wound density
ture indicating that among all the short- and long-lived between HaCaTs alone and those cocultured with fibroblasts,
ROS produced by CAP, H2O2 is one of the most stable oxi- and we observed that after CAP treatment, relative wound
dants produced by kINPen MED [41]. Based on this, we density was significantly increased in coculture compared to
quantified the H2O2 concentration (in μM) and found that that observed in HaCaT monoculture (Figure 2(c)). These
there is a linear increase in H2O2 concentration according results indicated that the CAP-mediated effect on wound
to treatment time (Fig S3). CAP did not significantly alter closure may differ between mono- and coculture. However,
metabolic activity after 10 and 30 s of treatment; however, a as coculture is more closely related to skin models due
decrease was observed in response to a 60 s of treatment after to the inclusion of both keratinocytes and fibroblasts, we
3 hr of incubation in all cell types (Figure 1(a)). Further assessed the signalling pathways that were active in each
significant reduction was not observed in the metabolic cell type to determine the stimulatory effect of CAP on
activity of HaCaTs and cocultured cells after 24 hr, but a coculture migration.
30 and 60 s treatment resulted in a significant reduction
in metabolic activity in GM Fbs after 24 hr (Figure 1(b)).
Based on this, we chose 10 and 30 s treatments for all 3.3. CAP Accelerates Migration through the Activation of the
subsequent experiments. YAP-CTGF-Cyr61 Axis. Quantitative real-time PCR revealed
 that CAP significantly stimulated YAP mRNA expression at
3.2. CAP Accelerates Cell Migration in Coculture. Migration 6 hr in GM Fbs (Figure 3(b)). CAP also tends to increase
and proliferation of keratinocytes and fibroblasts are key bio- YAP at late time points in GM Fbs (18 hr) and HaCaTs
logical processes that occur during various steps of wound (at 24 hr) (Figures 3(a) and 3(b)), and CAP minimally
healing, including wound repair and reepithelization. To induced CTGF and Cyr61 in HaCaTs at 3 hr (Figure 3(a)).
exclusively determine cell migration, all cell types were However, CAP increased CTGF and Cyr61 gene expression
starved prior to performing the scratch assay. Based on the to more than threefold in GM Fbs, and this peaked 18 hr after
Oxidative Medicine and Cellular Longevity 5

 Cell migration assay HaCaT Cell migration assay GM00637 Cell migration assay Coculture
 120 120
Relative wound density (%)

 Relative wound density (%)
 120

 Relative wound density (%)
 110 110 110
 100 100 100
 90 90 90
 80 80 80
 70 70 70
 60 60 60
 50 50 50
 40 40 40
 30 30 30
 20 20 20
 10 10 10
 0 0 0
 0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
 Time (hours) Time (hours) Time (hours)

 HaCaT control GM Fbs control Coculture control
 HaCaT 10 sec GM Fbs 10 sec Coculture 10 sec
 HaCaT 30 sec GM Fbs 30 sec Coculture 30 sec
 (a)
 CAP treatment HaCaT CAP treatment GM Fbs CAP treatment coculture
 120 120 120

 Relative wound density (%)
 Relative wound density (%)

 Relative wound density (%)

 100 ⁎⁎⁎⁎ ⁎ ⁎
 100 100
 ⁎⁎⁎⁎ ⁎
 80 80 80
 60 60 60
 40 40 40
 20 20 20
 0 0 0
 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr
 Control
 10 sec
 30 sec
 (b)
 Control 150 10 s CAP 30 s CAP
 150 150
 Relative wound density (%)

 Relative wound density (%)

 Relative wound density (%)

 ⁎
 100 100 ⁎⁎ 100
 ⁎

 50 50 50

 0 0 0
 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr

 Control HaCaT 10 s HaCaT 30 s HaCaT
 Control coculture 10 s coculture 30 s coculture
 (c)

Figure 2: CAP accelerates cell migration in coculture: (a) depicts the representative time course of wound closure in untreated or CAP-treated
HaCaTs, GM Fbs, and coculture, respectively, (upper row) up to 24 h. (b) shows the summation of relative wound density of untreated
and CAP-treated HaCaTs, GM Fbs, and coculture, respectively, (middle row) at specific time points (6, 12, 18, and 24 h). (c) details the
comparison of relative wound density between untreated and CAP-treated HaCaTs and coculture at 6, 12, 18, and 24 hr (lower row).
Data are presented as mean ± SD of three independent experiments performed in triplicate. Statistical comparisons were performed
using 2-way ANOVA.

treatment and then declined to near baseline by 24 hr not significantly altered in GM Fbs (Figure 4(a)); however,
(Figure 3(b)). there was a tendency for increased phosphorylation in
 Under serum-starved conditions, the ratio of phosphory- HaCaT cells (Figure 4(b)). We also observed an increase in
lated YAP compared to total YAP (pYAP/YAP) protein was total cellular CTGF (3 hr) and Cyr61 protein (3 hr and 6 hr)
6 Oxidative Medicine and Cellular Longevity

 YAP (HaCaT) 5 CTGF (HaCaT) Cyr61 (HaCaT)
 5 5
 Relative gene expression

 Relative gene expression

 Relative gene expression
 4 4 4
 (fold change)

 (fold change)

 (fold change)
 3 ⁎ ⁎⁎
 3 3
 ⁎ ⁎
 2 2 2

 1 1 1

 0 0 0
 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 (a)
 YAP (GM Fbs) CTGF (GM Fbs) Cyr61 (GM Fbs)
 ⁎
 5 5 5
 Relative gene expression

 Relative gene expression
 Relative gene expression

 4 0.06 4
 4
 (fold change)

 (fold change)
 (fold change)

 3 3 0.06 3
 ⁎
 ⁎ ⁎⁎
 2 2 2
 ⁎
 1 1 1

 0 0 0
 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 18 hr
 10 s 18 hr
 30 s 18 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 (b)

Figure 3: mRNA expression of HIPPO signalling effectors after CAP treatment: mRNA expression of YAP, CTGF, and Cyr61 in HaCaTs (a)
and GM Fbs (b) measured 3, 6, 18, and 24 hr after CAP treatment by qPCR and normalized to relative gene expression (ΔΔCT values on a log2
scale). The x-axis represents CAP treatment time and incubation time after plasma treatment. Data are represented as mean ± SD of either
three (a) or four (b) independent experiments. Statistical analysis was performed using unpaired t-test with Welch’s correction for
multiple comparisons, along with normalization to the untreated control.

levels after CAP treatment in GM Fbs (Figures 4(c) and 4(d)); NAC alone did not improve HaCaT cell migration. CAP treat-
however, unlike in GM Fbs, we could only detect a minor ment actually resulted in a modest acceleration in HaCaT cell
increase of CTGF in HaCaT cells upon CAP treatment migration that was significantly reduced after NAC adminis-
(Figure 4(e)) and Cyr61 protein expression was completely tration (Figure 5(b)). Similar results were obtained in cocul-
undetectable (data not shown). ture experiments, where NAC reduced CAP-mediated cell
 migration to ~40-50%. (Figure 5(c)).
3.4. Effect of NAC on Metabolic Activity and Cell Migration.
NAC is a well-known inhibitor of reactive species, and to 3.5. Effect of NAC on YAP-CTGF-Cyr61 Axis. We previously
determine the effect of CAP-produced ROS on cell signalling, observed a CAP-mediated activation of the YAP-CTGF-
cells were incubated with 2.5 mM NAC alone or in combina- Cyr61 signalling cascade. To confirm that the upregulation
tion with differential CAP-treated media for 24 hr. After of this signalling is mediated by CAP-induced reactive spe-
24 hr, NAC alone did not significantly alter the metabolic cies, total RNA was isolated from HaCaTs and GM Fbs after
activity of the cells. In HaCaTs, NAC significantly counter- incubation with NAC alone or with CAP-treated media. An
acted the 30 and 60 s CAP treatment effects. In GM Fbs, 18-hour incubation time in the presence of CAP-treated
NAC counteracted 10, 30, and 60 s CAP treatments and media was chosen for gene expression analysis. NAC in com-
increased the metabolic activity. In coculture, NAC signifi- bination with 10 and 30 s CAP-treated media significantly
cantly counteracted the effects of 30 s of CAP treatment and reduced the expression of CTGF (~2.4 fold) and Cyr61
increased the metabolic activity of these cells (Figure 5(a)). (~1.9 fold) in GM Fbs. YAP expression was also reduced;
Although metabolic activity was restored, we still sought to however, NAC alone did not significantly impact these sig-
determine if CAP-induced wound healing signalling is nalling molecules (Figure 6(b)). Similar results were obtained
affected after NAC administration. Therefore, cell migration in HaCaT cells that revealed that NAC significantly reduced
assays were performed using HaCaT monoculture and cocul- the CAP-mediated Cyr61 increase; however, YAP expression
ture. The cells were incubated with CAP-treated media with remained unaltered and CTGF expression was slightly but
or without 2.5 mM NAC. Interestingly, we observed that not significantly reduced (Figure 6(a)).
Oxidative Medicine and Cellular Longevity 7

 pYAP/YAP (GM Fbs) pYAP/YAP (HaCaT)
 3h 6h 24h
 5 5

 (relative fold change)

 (relative fold change)
 con 10 s 30 s con 10 s 30 s con 10 s 30 s 3h 6h 24 h
 4 con 10 s 30 s con 10 s 30 s con 10 s 30 s 4

 pYAP/tYAP

 pYAP/tYAP
 3 3
 pYAP
 2 pYAP 2
 1 1
 YAP
 0 YAP 0

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 (a) (b)
 3h 6h 24 h Cyr61 (GM Fbs)
 3h 6h 24 h 5 CTGF (GM Fbs) con 10 s 30 s con 10 s 30 s con 10 s 30 s

 Relative fold change
 5

 Relative fold change
 con 10 s 30 s con 10 s 30 s con 10 s 30 s 4 4
 3 Cyr61 3
 CTGF 2 2
 1 Actin 1
 Actin 0 0
 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr

 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 (c) (d)
 CTGF (HaCaT)
 3h 6h 24 h
 5

 Relative fold change
 con 10 s 30 s con 10 s 30 s con 10 s 30 s
 4
 3
 CTGF
 2
 1
 Actin
 0 con 3 hr
 10 s 3 hr
 30 s 3 hr
 con 6 hr
 10 s 6 hr
 30 s 6 hr
 con 24 hr
 10 s 24 hr
 30 s 24 hr
 (e)

Figure 4: Effect of CAP on the HIPPO signalling pathway: the relative phosphorylation level of YAP is shown in GM Fbs (a) and HaCaT cells
(b). CTGF and Cyr61 protein expression was enhanced after 10 and 30 s of CAP treatment in GM Fbs (c, d). CTGF expression slightly
increased after 10 and 30 s of CAP treatments in HaCaT cells after approximately 3 hours (e). Representative blots are shown. The x-axis
represents CAP treatment time and incubation time after plasma treatment. Data are represented as mean ± SD of at least three
independent experiments. Statistical analysis was performed using unpaired t-test with Welch’s correction for multiple comparisons, along
with normalization to the untreated control.

3.6. CAP Promotes Migration in Coculture through Secreted 4. Discussion
CTGF and Cyr61. CAP-induced migration was higher in
HaCaTs cocultured with GM Fbs than in monoculture alone During the past few years, plasma medicine has emerged as
(see Figure 2), indicating that CAP regulates the interaction an important field in biomedical science. Cold atmospheric
between these cells in coculture. We assumed that the major plasma generates a mixture of reactive oxygen and nitrogen
regulatory factors are CTGF and Cyr61 in coculture based on species, including singlet oxygen, atomic oxygen, hydroxyl
these data. We performed a cell migration assay to determine radicals, and hydrogen peroxide, as major (secondary) con-
the migratory behaviour of HaCaTs grown in media condi- tributors. Further, especially long-living species can be gener-
tioned from untreated or CAP-treated GM Fbs. Conditioned ated upon contact with liquids [42, 43]. Various studies have
medium harvested from CAP-treated GM Fbs at 18 hr signif- focused on the anti-inflammatory, antimicrobial, antineo-
icantly accelerated HaCaT cell migration compared to condi- plastic, and wound healing promoting properties of CAP
tioned medium from untreated GM Fbs and normal medium [13, 14, 44, 45]. In recent years, CAP has been used for the
(Figure 7(a)). We also analysed the GM-Fb-CM by ELISA, management of acute and chronic wounds in clinics; how-
and we observed an increase in secreted CTGF and Cyr61 ever, the mechanistic insight of CAP-mediated wound heal-
after CAP treatment (Figure 7(b)). ing responses is still not investigated enough. Here, not
 Recombinant CTGF and Cyr61 were also used to assess only did we demonstrate that CAP enhances wound healing
cell migration. HaCaT monoculture was treated with 10 in vitro, but we also provided evidence regarding the CAP-
and 50 ng/ml recombinant CTGF and 100 ng/ml and 1 μg/ml mediated induction of differentially regulated genes and the
of Cyr61. Treatment with 10 ng/ml CTGF and treatment paracrine effectors that promote wound healing via the
with 50 ng/ml significantly accelerated cell migration HIPPO pathway.
(Figure 7(c)). Treatment with either 100 ng/ml or 1 μg/ml Immortalized dermal keratinocytes and dermal fibro-
Cyr61 also accelerated cell migration at early time point blasts were used for our present study, as these cells are
(Figure 7(d)). primarily affected during skin/wound treatment. We also
8 Oxidative Medicine and Cellular Longevity

 HaCaT-24 hr GM Fbs- 24 hour Coculture-24 hr
 ⁎⁎⁎ ⁎
 ⁎ 200
 200 200 ⁎
 ⁎⁎ ⁎ ⁎⁎⁎
 % metabolic activity

 % metabolic activity

 % metabolic activity
 ⁎⁎⁎⁎
 150 150 ⁎⁎⁎⁎ 150

 100 100 100

 50 50 50

 0 0 0
 con
 NAC
 10 s
 10 s+NAC
 30 s
 30 s+NAC
 60 s
 60 s+NAC

 con
 NAC
 10 s
 10 s+NAC
 30 s
 30 s+NAC
 60 s
 60 s+NAC

 con
 NAC
 10 s
 10 s+NAC
 30 s
 30 s+NAC
 60 s
 60 s+NAC
 (a)

 HaCaT Coculture
 120
 120
 ⁎⁎ ⁎⁎⁎

 Relative wound density (%)
 ⁎⁎ 100 ⁎⁎⁎
 Relative wound density (%)

 100 ⁎⁎ ⁎⁎ ⁎
 ⁎⁎
 ⁎⁎ ⁎ 80 ⁎
 80
 ⁎⁎ ⁎⁎
 60 60

 40 40

 20 20

 0 0
 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr
 con 10 s+2.5 mM NAC
 NAC 30 s
 10 s 30 s+2.5 mM NAC
 (b) (c)

Figure 5: Effect of NAC on CAP-mediated wound healing and metabolic activity: metabolic activity of HaCaTs, GM Fbs, and coculture after
24 hr incubation with 2.5 mM NAC alone or in combination with CAP-treated media (a). Relative wound density of HaCaTs after treatment
with NAC alone or in combination with CAP-treated media at 6, 12, 18, and 24 hr (b). Relative wound density of coculture after treatment
with NAC alone or in combination with CAP-treated media at 6, 12, 18, and 24 hr (c). Data are presented as mean ± SD from at least three
independent experiments performed in triplicate (a–c). Statistical analysis was performed using either unpaired t-test with Welch’s correction
(a) or 2-way ANOVA (b, c).

chose the in vitro keratinocyte-fibroblast coculture model for ment [48, 49]. High amounts of ROS/RNS species can affect
a portion of our studies, as this model provides the ability to human cells and lead to cell death and reduction in cell via-
examine the influence of fibroblasts on cocultured keratino- bility and proliferation [50]. Indeed, we observed that a 60 s
cytes while omitting additional influences that are present CAP treatment significantly reduced cell viability in HaCaTs,
in intact skin. For our experiments, we used the comparison GM Fbs, and cocultures within 3 hr. Cell viability in GM Fbs
of monocultured fibroblasts and keratinocytes on the one significantly decreased by greater than 50% after 24 hr, and
hand side with the coculture of both cell types, in order to based on this, we excluded the 60 s treatment from our subse-
identify possible paracrine interactions of those cells after quent experiments. However, short- and long-lived radicals
treatment with CAP. These data are intended to support fur- also act as secondary messengers and contribute to wound
ther research in chronic wound healing—mostly to compare healing processes [51]. Therefore, the aim of this study was
data from diabetic patients. Therefore, it is important to to identify the stimulatory effects of CAP-generated reactive
understand how fibroblasts and keratinocytes influence each species on wound healing.
other to support wound healing after cold plasma treatment. The epidermal layer functions as a barrier to the outside
 H2O2 is an important second messenger that is generated of the skin, and upon damage, newly divided keratinocytes
by CAP which triggers the release of growth factors during migrate across the dermis [52]. Here, we optimized an
the wound healing process and has been shown to increase in vitro coculture model that incorporated keratinocytes
keratinocyte viability and migration in wound models [46]. and fibroblasts. Unlike in single cell studies, this model
Previous studies have revealed an increase in H2O2 in for migration studies using coculture provides the oppor-
response to increasing CAP treatments [47]. Intracellular tunity to monitor the influence of each cell type on the
ROS levels also increase in keratinocytes after plasma treat- other after CAP treatment. We observed that cells in
Oxidative Medicine and Cellular Longevity 9

 YAP (HaCaT) CTGF (HaCaT) Cyr61 (HaCaT)
 5 5 5
 Relative fold change

 Relative fold change

 Relative fold change
 4 4 4
 ⁎
 3 3 3
 ⁎ ⁎
 2 2 2

 1 1 1

 0 0 0
 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC

 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC

 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC
 (a)
 YAP (GM Fbs) CTGF (GM Fbs) Cyr61 (GM Fbs)
 ⁎
 5 ⁎
 5 5
 ⁎
 4
 Relative fold change

 Relative fold change

 Relative fold change
 4 4
 ⁎⁎⁎
 3 3 ⁎⁎ 3 ⁎

 2 2 2

 1 1 1

 0 0 0
 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC

 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC

 con

 NAC

 10 s

 10 s+NAC

 30 s

 30 s+NAC
 (b)

Figure 6: Effect of NAC on the YAP-CTGF-Cyr61 signalling cascade: mRNA expression of YAP, CTGF, and Cyr61 in HaCaTs (a) and GM
Fbs (b) measured 18 hr after CAP treatment by qPCR and normalized to relative gene expression (ΔΔCT values on a log2 scale). Data are
presented as mean ± SD from at least three independent experiments, and statistical analysis was performed using unpaired t-test with
Welch’s correction.

coculture closely mimic real skin, and this system also dis- cells. In accordance with this gene upregulation, CTGF and
played better CAP stimulation effects compared to those Cyr61 protein expression levels were also increased in GM
observed in the monocultures, indicating a crosstalk between Fbs. Upon phosphorylation of YAP, the HIPPO signalling
the plasma-treated cell types. Many growth factors and cyto- cascade is blocked and phosphorylated YAP is retained in
kines are released from fibroblasts after injury, and this could the cytoplasm. As YAP phosphorylation is increased after
provide an explanation for the acceleration of coculture CAP treatment in HaCaT cells, this could provide a possible
wound closure [53–55]. explanation for the minimal increase in CTGF protein and
 To more thoroughly assess the paracrine signalling axis, lack of Cyr61 detection in HaCaT cells; however, a significant
we examined the signalling pathways involved in tissue decrease in YAP phosphorylation was not observed in GM
homeostasis and regeneration. Emerging evidence suggests Fbs. Therefore, it is likely that the interaction of cytoplasmic
that the HIPPO pathway is regulated by a variety of signals, phosphorylated YAP with other signalling pathways (like
including mechanical stress, DNA damage, cellular stress, Wnt, Notch, and TGF-β) can also contribute to the increased
and oxidative stress [56]. Hence, we were interested in the translation of CTGF and Cyr61 in GM Fbs [58–62]. Further
HIPPO signalling pathway, as it plays an important role in studies are required to examine other upstream regulators
cell growth, proliferation, and tissue homeostasis and is that can induce CTGF and Cyr61 expressions.
under redox control [25, 57]. YAP, a multifunctional adapter The effect of NAC on restoring metabolic activity
protein that possesses one coactivator domain, is a nuclear upon reduction of oxidative stress has been studied previously
executer of the HIPPO signalling pathway. Based on our in other cell types [63]. In accordance with those results, in our
results, there was an increase in the mRNA expression of immortalized skin cells, we also found that NAC restored and
YAP in GM Fbs and HaCaT cells. We also detected a major enhanced metabolic activity after reduction of CAP-mediated
upregulation of the YAP target genes CTGF and Cyr61 that reactive species. This can be directly attributed to the ability
was present primarily in GM Fbs and minimally in HaCaT of NAC to scavenge excess ROS within the media. CAP
10 Oxidative Medicine and Cellular Longevity

 Scratch wound assay with GM-Fb CM
 ⁎
 ⁎
 ⁎ ⁎
 120 ⁎⁎ ⁎
 CTGF Cyr61
Relative wound density (%)

 2500 2500
 90

 Secreted CTGF (pg/ml)

 Secreted Cyr61 (pg/ml)
 2000 2000

 60 1500 1500

 1000 1000
 30
 500 500

 0 0 0
 6 hr 12 hr 18 hr 24 hr con 10 s CAP 30 s CAP con 10 s CAP 30 s CAP

 Control
 Untreated GM-Fb CM
 10 s GM-Fb CM
 30 s GM-Fb CM
 (a) (b)
 CTGF
 120
 ⁎
 ⁎⁎ 120 Cyr61
 Relative wound density (%)

 ⁎
 Relative wound density (%)

 90
 90
 ⁎⁎
 60 ⁎ ⁎⁎
 60

 30
 30

 0 0
 6 hr 12 hr 18 hr 24 hr 6 hr 12 hr 18 hr 24 hr
 Control Control
 10 ng/ml CTGF 100 ng/ml CYR61
 50 ng/ml CTGF 1 g/ml CYR61
 (c) (d)

Figure 7: Effect of CAP on paracrine signalling between keratinocytes and fibroblasts: HaCaT cells were incubated with conditioned media
harvested from untreated or 10 and 30 s CAP-treated GM Fbs. Relative wound density of HaCaT cells was monitored after 6, 12, 18,
and 24 hr (a). Secreted CTGF and Cyr61 (pg/ml) was analysed by ELISA (b). Relative wound density in HaCaT cells was monitored
after treatment with 10 and 50 ng/ml CTGF and 100 ng/ml and 1 μg/ml Cyr61 (c, d). Data are presented as mean ± SD from three
(a, c, d) or two (b) independent experiments, and statistical analysis was performed using 1- (b) or 2-way ANOVA (a, c, d).

treatment can also reduce intracellular GSH levels in caused downregulation of CTGF and Cyr61 expressions in
HaCaT cells as described previously by Dezest et al. [64]. GO orbital fibroblasts and dermal primary fibroblasts, respec-
As a cysteine-donating compound, NAC may act as a pre- tively [35, 36]. Therefore, we hypothesized that CAP could
cursor to replenish the depleted cellular GSH upon CAP stimulate wound healing by triggering crosstalk between
treatment, which in turn could induce the metabolic activ- fibroblasts and keratinocytes that is mediated by HIPPO
ity. Previous studies have also demonstrated the inhibitory signalling.
effect of NAC on epithelial cell migration [65, 66]. Similar Increasing numbers of studies are now focusing on the
to these studies, we found that CAP-induced cell migration paracrine signalling and interaction between keratinocytes
decreased by approximately 50% in response to 2.5 mM and fibroblasts in the context of wound repair [52, 67]. To
NAC treatment in HaCaT alone and in HaCaT cocultured test our hypothesis that CAP-induced stimulatory effects in
with GM Fbs. Treatment with NAC reduced the mRNA coculture are a result of paracrine crosstalk, we monitored
expression of CTGF and Cyr61 in GM Fbs and Cyr61 in the cell migration of HaCaTs that were incubated with condi-
HaCaTs, and this confirms the effect of CAP in promoting tioned media harvested from CAP-treated GM Fbs (GM-Fb-
the upregulation of this signalling pathway at the onset of CM) collected after 18 hr of CAP treatment. GM-Fb-CM was
wound healing. These findings are consistent with those of capable of accelerating HaCaT cell migration to a greater
previous studies which observed that antioxidant treatment degree compared to that of CM harvested from untreated
Oxidative Medicine and Cellular Longevity 11

 Wound healing

 Keratinocytes
 CTGF
 Cyr61
 Cyr61 Cyr61
 CTGF
 CTGF CTGF

 CAP
 (.OH, NO,
 Fibroblasts NOx,ONOO-
 ,OOH-)

 Cyr61
 YAP CTGF
 TEADs

Figure 8: Schematic of keratinocyte activation by CAP modulated fibroblasts: the primary event in this scheme is the generation of reactive
species in response to CAP treatment. These reactive species function as secondary messengers and stimulate the production of HIPPO
signalling effectors such as CTGF in HaCaT and CTGF and Cyr61 by dermal fibroblasts in the vicinity. These extracellular matrix proteins
that are released from fibroblasts in turn activate keratinocytes, accelerate migration, and promote wound healing.

Fbs and no CM at all. This phenomenon supports our short- and long-lived redox species. kINPen Med has already
hypothesis that CAP induces paracrine signalling between received approval for clinical applications in wound healing
these two cell types by releasing paracrine effectors and hence [12, 68]. Here, we revealed a mechanistic insight into wound
promoting keratinocyte migration. As expected, we could healing where CAP induced a YAP-CTGF-Cyr61 signalling
also detect an increase in secreted CTGF and Cyr61 expres- axis in dermal fibroblasts that was inhibited by antioxidant
sion in GM-Fb-CM treated with plasma. Additionally, HaCaT treatment. Although dermal keratinocytes could minimally
monocultures treated with recombinant CTGF and Cyr61 produce HIPPO downstream effector CTGF, it is the secre-
also exhibited improved cell migration. In particular, these tion of CTGF and Cyr61 from fibroblasts that influences
data indicate that CTGF and Cyr61 secreted from CAP- the activation of keratinocytes through paracrine signalling
treated fibroblasts act as paracrine effectors in wound healing. and improves cell migration upon wound onset.
These results further confirm our hypothesis that CAP pro-
motes a beneficial interaction between keratinocytes and
fibroblasts that ultimately contributes to wound healing. Abbreviations
 Here, for the first time, we demonstrated that CAP can
 CAP: Cold atmospheric plasma
stimulate a regenerative signalling pathway in dermal cells.
 CM: Conditioned medium
We observed that the cold plasma-mediated effect on skin
 CTGF: Connective Tissue Growth Factor
repair is mainly due to the activation of fibroblasts that when
 Cyr61: Cysteine-rich angiogenic protein 61
stimulated by CAP causes a paracrine stimulation of kerati-
 YAP: Yes-associated protein 1
nocytes by releasing extracellular proteins as indicated by
 TEAD: TEA domain family member
our coculture experiments. Together, these results confirm
 GM Fbs: GM00637 fibroblasts
the role of the HIPPO pathway in plasma-mediated wound
 NAC: N-Acetyl L cysteine
healing that is dominantly triggered by long-living ROS that
 TGF-β: Transforming growth factor β
are generated by cold plasma.
 RNS: Reactive nitrogen species
 An overview of the impact of CAP treatment on the
 ROS: Reactive oxygen species.
cellular redox signalling based on the present data is provided
in Figure 8.
 Data Availability
5. Conclusions
 The majority of the data can be found in the manuscript, and
CAP treatment results into complex crosstalk between redox further data used to support the findings of this study are
signalling pathways in mammalian cells via the impact of available from the corresponding authors upon request.
12 Oxidative Medicine and Cellular Longevity

Conflicts of Interest [9] S. Bekeschus, A. Schmidt, L. Bethge et al., “Redox stimulation
 of human THP-1 monocytes in response to cold physical
The authors declare no conflict of interest regarding the plasma,” Oxidative Medicine and Cellular Longevity, vol. 2016,
publication of this paper. Article ID 5910695, 11 pages, 2016.
 [10] S. Arndt, M. Landthaler, J. L. Zimmermann et al., “Effects of
 cold atmospheric plasma (CAP) on ß-defensins, inflammatory
Acknowledgments cytokines, and apoptosis-related molecules in keratinocytes
 in vitro and in vivo,” PLoS One, vol. 10, no. 3, article
The authors acknowledge the technical assistance of Liane e0120041, 2015.
Kantz. This work was partly supported by TBI-V-1-234- [11] A. Kramer, S. Bekeschus, R. Matthes et al., “Cold physical
VBW-081 from the Federal Government of Mecklenburg, plasmas in the field of hygiene—relevance, significance, and
Western Pomerania. future applications,” Plasma Processes and Polymers, vol. 12,
 no. 12, pp. 1410–1422, 2015.
 [12] A. Schmidt, S. Bekeschus, K. Wende, B. Vollmar, and T. von
Supplementary Materials Woedtke, “A cold plasma jet accelerates wound healing in a
 murine model of full-thickness skin wounds,” Experimental
Figure S1: summary of cell count in monoculture and cocul-
 Dermatology, vol. 26, no. 2, pp. 156–162, 2017.
ture. Figure S2: forward and reverse primer sequences used
 [13] S. Arndt, A. Schmidt, S. Karrer, and T. von Woedtke, “Compar-
in real-time PCR for YAP, Cyr61, and CTGF. Figure S3:
 ing two different plasma devices kINPen and Adtec SteriPlas
Amplex Red assay was performed to quantitatively measure regarding their molecular and cellular effects on wound heal-
the production of H2O2 in response to CAP treatment. ing,” Clinical Plasma Medicine, vol. 9, pp. 24–33, 2018.
Figure S4: wound healing images at times 0, 12, and 24 hr
 [14] A. Schmidt, T. von Woedtke, B. Vollmar, S. Hasse, and
in untreated and CAP-treated coculture. Figure S5: wound S. Bekeschus, “Nrf2 signaling and inflammation are key events
healing images at times 0, 12, and 24 hr in untreated and in physical plasma-spurred wound healing,” Theranostics,
CAP-treated HaCaT cells. Figure S6: wound healing images vol. 9, no. 4, pp. 1066–1084, 2019.
at times 0, 12, and 24 hr in untreated and CAP-treated GM [15] G. Halder and R. L. Johnson, “Hippo signaling: growth control
Fbs. (Supplementary Materials) and beyond,” Development, vol. 138, no. 1, pp. 9–22, 2011.
 [16] D. Pan, “The hippo signaling pathway in development and
References cancer,” Developmental Cell, vol. 19, no. 4, pp. 491–505, 2010.
 [17] F. X. Yu and K. L. Guan, “The Hippo pathway: regulators and
 [1] T. von Woedtke, A. Schmidt, S. Bekeschus, K. Wende, and
 regulations,” Genes & Development, vol. 27, no. 4, pp. 355–371,
 K.-D. Weltmann, “Plasma medicine: a field of applied redox
 2013.
 biology,” In Vivo, vol. 33, no. 4, pp. 1011–1026, 2019.
 [18] W. Hong and K. L. Guan, “The YAP and TAZ transcription
 [2] L. E. Reynolds, F. J. Conti, M. Lucas et al., “Accelerated
 co-activators: key downstream effectors of the mammalian
 re-epithelialization in β3-integrin-deficient- mice is associated
 Hippo pathway,” Seminars in Cell & Developmental Biology,
 with enhanced TGF-β1 signaling,” Nature Medicine, vol. 11,
 vol. 23, no. 7, pp. 785–793, 2012.
 no. 2, pp. 167–174, 2005.
 [3] A. Schmidt, S. Dietrich, A. Steuer et al., “Non-thermal plasma [19] M. J. Lee, M. R. Byun, M. Furutani-Seiki, J. H. Hong, and H. S.
 activates human keratinocytes by stimulation of antioxidant Jung, “YAP and TAZ regulate skin wound healing,” Journal of
 and phase II pathways,” Journal of Biological Chemistry, Investigative Dermatology, vol. 134, no. 2, pp. 518–525, 2014.
 vol. 290, no. 11, pp. 6731–6750, 2015. [20] X. Luo, Y. Liu, W. Feng et al., “NUP37, a positive regulator of
 [4] R. Melchionna, G. Bellavia, M. Romani et al., “C/EBPγ regu- YAP/TEAD signaling, promotes the progression of hepatocel-
 lates wound repair and EGF receptor signaling,” Journal of lular carcinoma,” Oncotarget, vol. 8, no. 58, pp. 98004–98013,
 Investigative Dermatology, vol. 132, no. 7, pp. 1908–1917, 2017.
 2012. [21] E. R. Barry and F. D. Camargo, “The Hippo superhighway:
 [5] H. R. Rezvani, N. Ali, M. Serrano-Sanchez et al., “Loss of epi- signaling crossroads converging on the Hippo/Yap pathway
 dermal hypoxia-inducible factor-1α accelerates epidermal in stem cells and development,” Current Opinion in Cell
 aging and affects re-epithelialization in human and mouse,” Biology, vol. 25, no. 2, pp. 247–253, 2013.
 Journal of Cell Science, vol. 124, no. 24, pp. 4172–4183, 2011. [22] D. P. Del Re, Y. Yang, N. Nakano et al., “Yes-associated protein
 [6] C. Dunnill, T. Patton, J. Brennan et al., “Reactive oxygen spe- isoform 1 (Yap1) promotes cardiomyocyte survival and
 cies (ROS) and wound healing: the functional role of ROS growth to protect against myocardial ischemic injury,” Journal
 and emerging ROS-modulating technologies for augmentation of Biological Chemistry, vol. 288, no. 6, pp. 3977–3988, 2013.
 of the healing process,” International Wound Journal, vol. 14, [23] B. Zhao, L. Li, Q. Lu et al., “Angiomotin is a novel Hippo path-
 no. 1, pp. 89–96, 2017. way component that inhibits YAP oncoprotein,” Genes &
 [7] L. Bundscherer, K. Wende, K. Ottmüller et al., “Impact of non- Development, vol. 25, no. 1, pp. 51–63, 2011.
 thermal plasma treatment on MAPK signaling pathways of [24] B. Zhao, X. Wei, W. Li et al., “Inactivation of YAP oncoprotein
 human immune cell lines,” Immunobiology, vol. 218, no. 10, by the Hippo pathway is involved in cell contact inhibition and
 pp. 1248–1255, 2013. tissue growth control,” Genes & Development, vol. 21, no. 21,
 [8] K. Wende, S. Straßenburg, B. Haertel et al., “Atmospheric pp. 2747–2761, 2007.
 pressure plasma jet treatment evokes transient oxidative stress [25] Y. Wang, A. Yu, and F. X. Yu, “The Hippo pathway in tissue
 in HaCaT keratinocytes and influences cell physiology,” Cell homeostasis and regeneration,” Protein & Cell, vol. 8, no. 5,
 Biology International, vol. 38, no. 4, pp. 412–425, 2014. pp. 349–359, 2017.
Oxidative Medicine and Cellular Longevity 13

[26] C. C. Chen, F. E. Mo, and L. F. Lau, “The angiogenic factor [41] K. Wende, P. Williams, J. Dalluge et al., “Identification of the
 Cyr61 activates a genetic program for wound healing in biologically active liquid chemistry induced by a nonthermal
 human skin fibroblasts,” Journal of Biological Chemistry, atmospheric pressure plasma jet,” Biointerphases, vol. 10,
 vol. 276, no. 50, pp. 47329–47337, 2001. no. 2, article 029518, 2015.
[27] B. Zhao, X. Ye, J. Yu et al., “TEAD mediates YAP-dependent [42] P. J. Bruggeman, M. J. Kushner, B. R. Locke et al., “Plasma–
 gene induction and growth control,” Genes & Development, liquid interactions: a review and roadmap,” Plasma Sources
 vol. 22, no. 14, pp. 1962–1971, 2008. Science and Technology, vol. 25, no. 5, article 053002, 2016.
[28] C.-C. Chen and L. F. Lau, “Functions and mechanisms of [43] K. Wende, T. von Woedtke, K. D. Weltmann, and
 action of CCN matricellular proteins,” The International Jour- S. Bekeschus, “Chemistry and biochemistry of cold physical
 nal of Biochemistry & Cell Biology, vol. 41, no. 4, pp. 771–783, plasma derived reactive species in liquids,” Biological Chemis-
 2009. try, vol. 400, no. 1, pp. 19–38, 2018.
[29] W. H. Fan, M. Pech, and M. J. Karnovsky, “Connective tissue [44] S. Arndt, E. Wacker, Y. F. Li et al., “Cold atmospheric plasma, a
 growth factor (CTGF) stimulates vascular smooth muscle cell new strategy to induce senescence in melanoma cells,” Experi-
 growth and migration in vitro,” European Journal of Cell Biol- mental Dermatology, vol. 22, no. 4, pp. 284–289, 2013.
 ogy, vol. 79, no. 12, pp. 915–923, 2000. [45] A. M. Hirst, F. M. Frame, M. Arya, N. J. Maitland, and
[30] Z. Yang, Z. Sun, H. Liu et al., “Connective tissue growth factor D. O'Connell, “Low temperature plasmas as emerging cancer
 stimulates the proliferation, migration and differentiation of therapeutics: the state of play and thoughts for the future,”
 lung fibroblasts during paraquat-induced pulmonary fibrosis,” Tumour Biology, vol. 37, no. 6, pp. 7021–7031, 2016.
 Molecular Medicine Reports, vol. 12, no. 1, pp. 1091–1097, [46] A. E. Loo, Y. T. Wong, R. Ho et al., “Effects of hydrogen perox-
 2015. ide on wound healing in mice in relation to oxidative damage,”
[31] D. R. Brigstock, “Regulation of angiogenesis and endothelial PLoS One, vol. 7, no. 11, article e49215, 2012.
 cell function by connective tissue growth factor (CTGF) and [47] R. K. Gandhirajan, K. Rödder, Y. Bodnar et al., “Cytochrome C
 cysteine-rich 61 (CYR61),” Angiogenesis, vol. 5, no. 3, oxidase inhibition and cold plasma-derived oxidants synergize
 pp. 153–165, 2002. in melanoma cell death induction,” Scientific Reports, vol. 8,
[32] M. P. Alfaro, D. L. Deskins, M. Wallus et al., “A physiological no. 1, article 12734, 2018.
 role for connective tissue growth factor in early wound heal- [48] A. Schmidt, S. Bekeschus, K. Jarick, S. Hasse, T. von Woedtke,
 ing,” Laboratory Investigation, vol. 93, no. 1, pp. 81–95, 2013. and K. Wende, “Cold physical plasma modulates p53 and
[33] J. I. Jun and L. F. Lau, “The matricellular protein CCN1 mitogen-activated protein kinase signaling in keratinocytes,”
 induces fibroblast senescence and restricts fibrosis in cutane- Oxidative Medicine and Cellular Longevity, vol. 2019, Article
 ous wound healing,” Nature Cell Biology, vol. 12, no. 7, ID 7017363, 16 pages, 2019.
 pp. 676–685, 2010. [49] A. Schmidt, T. von Woedtke, and S. Bekeschus, “Periodic
[34] J. M. Schober, L. F. Lau, T. P. Ugarova, and S. C.-T. Lam, exposure of keratinocytes to cold physical plasma: an in vitro
 “Identification of a novel integrin αMβ2 binding site in CCN1 model for redox-related diseases of the skin,” Oxidative Med-
 (CYR61), a matricellular protein expressed in healing wounds icine and Cellular Longevity, vol. 2016, Article ID 9816072,
 and atherosclerotic lesions,” Journal of Biological Chemistry, 17 pages, 2016.
 vol. 278, no. 28, pp. 25808–25815, 2003. [50] A. K. Kaushik, S. K. Vareed, S. Basu et al., “Metabolomic profil-
[35] C. C. Tsai, S. B. Wu, H. C. Kau, and Y. H. Wei, “Essential role ing identifies biochemical pathways associated with castration-
 of connective tissue growth factor (CTGF) in transforming resistant prostate cancer,” Journal of Proteome Research, vol. 13,
 growth factor-β1 (TGF-β1)-induced myofibroblast transdif- no. 2, pp. 1088–1100, 2014.
 ferentiation from Graves' orbital fibroblasts,” Scientific Reports, [51] A. Soneja, M. Drews, and T. Malinski, “Role of nitric oxide,
 vol. 8, no. 1, p. 7276, 2018. nitroxidative and oxidative stress in wound healing,” Pharma-
[36] Z. Qin, P. Robichaud, T. He, G. J. Fisher, J. J. Voorhees, and cological Reports, vol. 57, pp. 108–119, 2005.
 T. Quan, “Oxidant exposure induces cysteine-rich protein 61 [52] G. Seo, C. Hyun, D. Koh et al., “A novel synthetic
 (CCN1) via c-Jun/AP-1 to reduce collagen expression in material, BMM, accelerates wound repair by stimulating
 human dermal fibroblasts,” PLoS One, vol. 9, no. 12, article re-epithelialization and fibroblast activation,” International
 e115402, 2014. Journal of Molecular Sciences, vol. 19, no. 4, p. 1164,
[37] S. Bekeschus, A. Schmidt, K. D. Weltmann, and T. von 2018.
 Woedtke, “The plasma jet kINPen – A powerful tool for [53] N. Maas-Szabowski, H. J. Stark, and N. E. Fusenig, “Kerati-
 wound healing,” Clinical Plasma Medicine, vol. 4, no. 1, nocyte growth regulation in defined organotypic cultures
 pp. 19–28, 2016. through IL-1-induced keratinocyte growth factor expression
[38] A. Schmidt, T. Woedtke, J. Stenzel et al., “One year follow-up in resting fibroblasts,” Journal of Investigative Dermatology,
 risk assessment in SKH-1 mice and wounds treated with an vol. 114, no. 6, pp. 1075–1084, 2000.
 argon plasma jet,” International Journal of Molecular Sciences, [54] N. Maas-Szabowski, A. Szabowski, H. J. Stark et al., “Organo-
 vol. 18, no. 4, p. 868, 2017. typic cocultures with genetically modified mouse fibroblasts as
[39] K. Wende, S. Bekeschus, A. Schmidt et al., “Risk assessment a tool to dissect molecular mechanisms regulating keratinocyte
 of a cold argon plasma jet in respect to its mutagenicity,” growth and differentiation,” Journal of Investigative Dermatol-
 Mutation Research/Genetic Toxicology and Environmental ogy, vol. 116, no. 5, pp. 816–820, 2001.
 Mutagenesis, vol. 798-799, pp. 48–54, 2016. [55] Z. Wang, Y. Wang, F. Farhangfar, M. Zimmer, and Y. Zhang,
[40] K. J. Livak and T. D. Schmittgen, “Analysis of relative gene “Enhanced keratinocyte proliferation and migration in
 expression data using real-time quantitative PCR and the co-culture with fibroblasts,” PLoS One, vol. 7, no. 7, article
 2−ΔΔCT method,” Methods, vol. 25, no. 4, pp. 402–408, 2001. e40951, 2012.
You can also read